Mechanism of Formation of Copper(II) Chloro Complexes Revealed by

Green State University, Bowling Green, Ohio 43403, United States. J. Phys. Chem. B , 2015, 119 (28), pp 8754–8763. DOI: 10.1021/acs.jpcb.5b03889...
0 downloads 5 Views 1MB Size
Subscriber access provided by UNIV OF MISSISSIPPI

Article

Mechanism of Formation of Copper(II) Chlorocomplexes Revealed by Transient Absorption Spectroscopy and DFT/TDDFT-Calculations. Andrey S Mereshchenko, Pavel K Olshin, Kanykey E. Karabaeva, Maxim S. Panov, Robert Marshall Wilson, Vladimir Alexeevich Kochemirovsky, Mikhail Yu. Skripkin, Yury S. Tveryanovich, and Alexander N. Tarnovsky J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.5b03889 • Publication Date (Web): 16 Jun 2015 Downloaded from http://pubs.acs.org on June 20, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry B is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Mechanism of Formation of Copper(II) Chlorocomplexes Revealed by Transient Absorption Spectroscopy and DFT/TDDFT-Calculations Andrey S. Mereshchenko,a,* Pavel K. Olshin,aKanykey E Karabaeva,b Maxim S. Panov,aR. Marshal Wilson,bVladimir A. Kochemirovsky,a Mikhail Yu. Skripkin,aYury S. Tveryanovich,a and Alexander N. Tarnovskyb a

Institute of Chemistry, Saint-Petersburg State University, 198504 Saint-Petersburg, Russian Federation.

b

Department of Chemistry and the Center for Photochemical Sciences, Bowling Green State University, Bowling Green, Ohio 43403, USA.

Keywords Copper(II), chlorocomplexes, pump-probe, complex formation, ligand exchange, ultrafast photochemistry, solution, dynamics.

Abstract

Copper(II) complexes are extremely labile with typical ligand exchange rate constants of the order of 106-109 M-1 s-1. As the result, it is often difficult to identify the actual formation

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 46

mechanism of these complexes. In this work, using UV-vis transient absorption when probing in a broad time range (20 ps- 8 µs) in conjunction with DFT/TDDFT calculations, we studied the dynamics and underlying reaction mechanisms of the formation of extremely labile copper(II) CuCl42-chlorocomplexes from copper(II) CuCl3-trichlorocomplexes and chloride ions. These two species, produced via photochemical dissociation of CuCl42- upon 420-nm excitation into the ligand-to-metal-charge-transfer electronic state, are found to recombine into parent complexes with bimolecular rate constants of (9.0 ± 0.1)·107 and (5.3 ± 0.4)·108 M-1 s-1 in acetonitrile and dichloromethane, respectively. In dichloromethane, recombination occurs via a simple one-step addition. In acetonitrile, where [CuCl3]- reacts with the solvent to form a [CuCl3CH3CN]complex in less than 20 ps, recombination takes place via ligand exchange described by the associative interchange mechanism that involves a [CuCl4CH3CN]2- intermediate. In both solvents, the recombination reaction is potential-energy controlled.

I.

Introduction

It is well known that copper plays an important role in metabolism of living organisms, for example, plastocyanin takes part as an electron donor species in the photosynthesis of the photosystem I in plants and bacteria.1 Electron and energy transport processes of this and several other copper-containing proteins have been the focus of many recent studies.2-4 In general, these transfer properties are associated with the ability of a copper ion to change its oxidation number reversibly.3 However, plastocyanin is a rather complicated system for studying the photoinduced transformation of the copper proteins. On the other hand, copper(II) chlorocomplexes are known as model compounds to understand the photochemistry of more complex copper complexes, and potentially, they can be useful for better understanding of copper proteins.5,6

ACS Paragon Plus Environment

2

Page 3 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

One of the most important aspects of coordination chemistry is formation dynamics of metal complexes, which usually involves ligand substitution (or exchange) processes. This means that it is necessary to know the ligand exchange mechanism of metal complexes in various environments to better understand the reactivity of metal ions in biology and biochemistry. The ligand exchange processes have been best studied in aqueous solutions,9,11,12,13, but also there is a number of studies describing ligand exchange processes in non-aqueous solutions7-9 and polymers.10 Substitution processes in non-aqueous media are of interest because of synthetic applications. .Ligand substitution kinetics of transition metals, e. g., manganese,14 chromium,11 nickel,12 cadmium,15 and others16 attract a great deal of attention and have been actively studied. Also, photoinduced structural rearrangements of metal complexes have been of interest.17 Depending on ligand exchange reaction rates, metal complexes are classified into two categories:18 inert complexes, where ligand exchange is slow (rate constants, < 101 M-1 s-1), and labile complexes characterized by high ligand exchange rate constants (> 102 M-1 s-1). Langford and Gray19 proposed to classify the mechanisms of ligand exchange reactions into associative, dissociative and interchange mechanisms. The associative mechanism (A) involves the reaction intermediate species with the coordination number larger than that of the initial complex due to association with the incoming ligand. On the contrary, the dissociative mechanism (D) involves dissociation of a ligand with formation of the intermediate species, which coordination number is smaller than that of the initial complex. The case between these two extremes is the interchange mechanism (I) that usually proceeds through a manifold of transition states separating intermediate complexes, which cannot be kinetically detected. Interchange mechanisms range from associative interchange (IA), in which the bond-making is of greater importance, to interchange (I), in which bond-making and bond-breaking processes are equally important, and

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 46

to dissociative interchange in which bond-breaking is predominantly important (ID).7 The A, IA, and ID mechanisms are common for four8- ,five7-, and six9- coordinated copper(II) complexes, respectively. The experimental kinetic data for copper complexes are extremely limited despite the fact that the thermodynamics of copper(II) ion complexation is studied rather well.20–24 The kinetics of a large number of stable, inert complexes of the transition metals can be studied by spectrophotometric methods.15 But, due to combination of Jahn-Taller distortion and fast reorientation of the distortion axis, copper(II) complexes are extremely labile with typical ligandexchange rate constants of the order of 106 - 109 M-1 s-1

7,9,25–27

and, in consequence,

conventional spectrophotometric methods fail to resolve the course of structural changes from reagents to products in copper(II) ligand exchange reactions. High-pressure NMR spectroscopy is one of the most informative and useful methods for describing the exchange mechanism of labile complexes.28,29 The results obtained by NMR studies can be supported by quantum chemical calculations.11 E.g., DFT, as one of the most reliable low-cost computation methods,30 allows to evaluate the thermodynamics and structural changes during the course of the reaction, and thus provides enough information to interpret the reaction mechanism. Another useful experimental for measuring ligand exchange rates and establishing the involved reaction mechanism is temperature-jump (T-jump) time-resolved transient mid-IR absorbance spectroscopy.

13

This method is based on local ultrafast heating of

the investigated substances, resulting in the equilibrium shift from initial complexes to product complexes. The T-jump method has a significantly better time resolution (5 ps)31 than NMR spectroscopy (several ns), and, therefore, is more suitable for studying ligand exchange reactions of extremely labile complexes such as copper(II) compounds. In this work, we demonstrate the

ACS Paragon Plus Environment

4

Page 5 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

power of studying the formation dynamics of very labile complexes by using picosecond and nanosecond UV-vis transient absorption spectroscopies in conjunction with DFT calculations. For this study, we selected copper(II) tetrachlorocomplexes for several reasons. To begin with, photolysis of [CuCl4]2- results in ionic dissociation without any reduction of copper(II) ions,32,33 so that the subsequent recombination of ionic fragment species can be exclusively monitored using transient absorption spectroscopy. Next, this complex has a tractable size for quantumchemical calculations. Also, [CuCl4]2- is soluble in both polar and nonpolar solvents, which makes it possible to study ligand substitution by solvents with different donor properties, i.e. different abilities of the solvent to coordinate to the metal ion (for example, dichloromethane and acetonitrile, which are characterized by donor numbers of 1 and 14.1,34 respectively). A small size of a chloride ligand minimizes the role of steric effects in these reactions. Herein, the photochemistry of [CuCl4]2- complexes is studied in two solvents of different donor properties, namely, dichloromethane and acetonitrile, by means of UV-vis transient absorption spectroscopy with probing from several picoseconds to several microseconds, and functional theory (DFT) and time-dependent density functional theory (TDDFT) calculations. We focus on the recombination reactions of the ionic dissociation products assigned, based on this work, to ligand exchange and ligand addition in acetonitrile and dichloromethane, respectively.

II.

Experimental and computational methods

Copper(II) perchlorate hexahydrate (98%), tetraethylammonium chloride (> 98%), dichloromethane (HPLC grade, anhydrous), and acetonitrile (> 99.5 %, anhydrous) were purchased from Sigma Aldrich. Tetraethylammonium chloride was dried in vacuum oven at

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 46

1100C for 10 hours. All experiments were performed in acetonitrile and dichloromethane. The series of solutions were prepared for nanosecond transient absorption experiments. In these solutions the concentration of copper(II) ions in a form of copper(II) perchlorate was kept constant at 0.1mM in both solvents. On the other hand, the concentration of chloride ions in a form of tetraethilammonium chloride was varied to be 10, 20, 50, 100 mM in acetonitrile and 1, 2, 5, 20 mM in dichloromethane. Ion ratio affects the ligand substitution process, therefore, we change the mole ratio of Cu(II):Cl- to study the kinetics of ligand exchange mechanism. For picosecond experiments we prepared the separate solutions in acetonitrile and dichloromethane in which the concentrations of Cu(II) and Cl- were 0.5 and 100 mM, respectively. UV-vis-NIRIR absorption spectra of the solutions were measured using a Varian Cary 50 UV-Vis spectrophotometer. All spectroscopic measurements were performed at 21oC. Picoseconds transient absorption spectra were measured using the previously described set-up based on a regeneratively amplified Ti: sapphire laser system (800 nm, 100 fs, and 1 kHz).32 An optical parametric amplifier (OPA) was used to generate 420-nm excitation (“pump”) pulses with the energy of 9 µJ pulse-1. White-light continuum probe pulses in the 340 - 760 spectral range were generated by focusing the 800-nm light into a 3-mm thick CaF2 window. The probe light was focused onto the sample to a 160-µm diameter spot and overlapped at an angle of 6° with the pump light focused to a 460-µm diameter spot. Transient absorption kinetic traces were measured simultaneously within 274-nm spectral intervals using a spectrograph/dual-diode (analyzing/reference) array detector. The polarization of the pump and probe light was set at the magic angle (54.7o) with respect to each other. The solutions were circulated through a 2-mm Spectrosil UV quartz flow cell.

ACS Paragon Plus Environment

6

Page 7 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Nanosecond transient absorption spectra were measured using a Proteus spectrometer (Ultrafast Systems) coupled with a nanosecond Nd:YAG laser/OPO system (Opotek, Vibrant LD 355 II, 410 - 2400 nm, 5 ns, and 10 Hz). The Proteus spectrometer was equipped with a 150 W Xe arc lamp (Newport) used as a probe-light source, a monochromator (Bruker Optics), and a Si photodiode (Thorlabs, DET 10A). The Nd:YAG laser/OPO system was set up to generate 420nm excitation (“pump”) pulses with the energy of 3 mJ pulse-1. Transient absorption signals were measured in the transverse geometry in which the probe was made to pass through a 1-cm pathlength quartz cell filled with the sample solution. The 1024 kinetic traces at a single probe wavelength were averaged and then analyzed using Origin 9.0 software. The geometries of the following species were optimized using DFT (B3LYP/6-31G(d)/PCM level of theory): [CuCl4]2- and [CuCl3]- (in acetonitrile and dichloromethane), CH3CN, [CuCl3CH3CN]-, and [CuCl4CH3CN]2- (acetonitrile only), and (NEt4)2CuCl4 and NEt4CuCl3 (dichloromethane only). Polarizable continuum model (PCM) was implemented to simulate solvent media, acetonitrile and dichloromethane. For the optimized geometries of the aforementioned Cu(II) species, vertical excitation transition (VET) energies were calculated using TDDFT (B3LYP/6-31G(d)). To find the energy profile of the reactions between copper(II) triclorocomplexes and chloride ions, relaxed redundant scans along Cu−Cl and Cu−N bonds were performed for [CuCl4]2-, [CuCl3CH3CN]-, and [CuCl4CH3CN]2-complexes. A relaxed redundant scan along a Cu(CuCl3-)−Cl(CH2Cl2) bond was performed to locate the minimum energy geometry of the [CuCl3CH2Cl2]- complex. All calculations were performed using Gaussian 09 software package.35

III.

Results

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry

Steady-state absorption spectroscopy In this study, [CuCl4]2- complexes were obtained by addition of excess of tetraethylammonium chloride to copper(II) perchlorate solutions. Acetonitrile and dichloromethane solutions containing 0.5 mM Cu(ClO4)2 and 100 mM NEt4Cl have all copper(II) ions existing in the form of [CuCl4]2-,32,33 exhibiting steady-state absorption spectra due to transitions in this species, Fig. 1. The intense UV-vis absorption bands of [CuCl4]2- correspond to ligand-to-metal chargetransfer (LMCT) transitions.20,21,32,36,37 The shape and position of the LMCT bands is observed to be similar for acetonitrile and dichloromethane with only a small spectral shift of the absorption maxima. Steady-state absorption spectra of other solutions used in the nanosecond experiments demonstrate that the majority of copper(II) exists in the form of [CuCl4]2- complexes (Fig. 1S and 2S, Supporting Information).

-1

6000

4000

-1

ε, M cm

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 46

2000

0

250 300 350 400 450 500 550 600

λ, nm

Figure. 1. Steady-state absorption spectra of [CuCl4]2- complexes in acetonitrile (solid line) and dichloromethane (dashed line).

Picosecond transient absorption spectroscopy

ACS Paragon Plus Environment

8

Page 9 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

For the picosecond transient absorption (∆A) measurements, we used solutions containing 0.5 mMCu(ClO4)2 and 100mM NEt4Cl in acetonitrile and dichloromethane. Upon 420-nm excitation, [CuCl4]2-, the predominant form in these solutions, is promoted into the LMCT excited state. Then, the minor fraction of the LMCT-excited complex undergoes ionic dissociation whereas the major one returnes back to the ground state via internal conversion, which involves the relaxation to the vibrationally hot ground and ligand field exited states. The lifetime of LMCT exited state is less than 500 fs and the short-time transient absorption spectra (0.5 - 20 ps) of [CuCl4]2-are dominated by the absorption of ligand-field excited states.32,33 The ligand-field excited states have a lifetime less than 5 ps, and therefore, they fully decays at 20 ps after excitation. Starting from 20 ps, the transient absorption spectra do not change in the both solvents investigated over the full delay time range (1 ns) employed in these experiments, Fig. 2. In acetonitrile, these ∆A spectra consisting of the 350-380 nm and 465-nm ∆A bands (induced absorption) and the negative bleach centered at 405 nm were previously assigned32,33 to the formation of copper(II) trichlorocomplexes, the product of ionic dissociation: 420nm

[Cu IICl4 ]2 −  →[Cu IICl3 ]− + Cl−

(1)

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry

0.0004

∆A

0.0000 20 ps 50 ps 100 ps 200 ps 500 ps 1 ns

A

-0.0004 -0.0008 0.0004

∆A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 46

0.0000

20 ps 50 ps 100 ps 200 ps 500 ps 1 ns

B

-0.0004 350

400

450

500

550

600

Wavelength, nm Figure. 2. Picosecond transient absorption spectra of [CuCl4]2- complexes in acetonitrile (panel A) and dichloromethane (panel B) following 420-nm excitation (0.5 mMCu(ClO4)2 - 100 mM NEt4Cl). Delay times between 420-nm excitation and probe pulses are shown in the legends. The ∆A spectra of [CuCl4]2-in dichloromethane in the 20 ps – 1 ns time range are similar to those in acetonitrile and, therefore, can also be assigned to the formation of copper(II) trichlorocomplexes.

Nanosecond transient absorption spectroscopy For the nanosecond transient absorption measurements, we used solutions containing 0.1 mM Cu(ClO4)2 and xmM NEt4Cl (where x varies from 10 to 100 in acetonitrile and from 1 to 20 in dichloromethane), Fig. 3-5. At 50 ns after 420-nm excitation of acetonitrile and dichloromethane solutions, the ∆A spectra are found to be similar, consisting of the 350-380 nm feature, the 465-

ACS Paragon Plus Environment

10

Page 11 of 46

nm ∆A band (shifted to 470 nm in dichloromethane), and the 405-nm negative bleach. The induced absorption bands decays to zero concurrently with recovery of the ground-state bleach and without any spectral reshaping. The decay of the ∆A bands is single-exponential, ∆A(t)= ∆A0·e-kobst (Fig. 4 and 5), where the pseudo-first-order rate constant (kobs) is found to linearly depend on the concentration of NEt4Cl, kobs=kq·[NEt4Cl], Fig. 6. Based on the kobs values resulting from these single-exponential fits, the bimolecular rate constants, kq, are evaluated to be (9.0 ± 0.1)·107 and (5.3 ± 0.4)·108 M-1 s-1 in acetonitrile and dichloromethane, respectively. The nanosecond and picoseconds ∆A spectra upon 420-nm excitation of[CuCl4]2- are identical, which suggests the formation of the copper(II) trichlorocomplex product in both experiments. The complete decay of the ∆A bands accompanied by the ground-state bleach recovery is assigned to recombination of copper(II) trichlorocomplexes and chloride ions into parent [CuCl4]2complexes.

0.002

∆A

0.001 0.000

50 ns 100 ns 200 ns 500 ns 1 µs 2 µs 5 µs

-0.001 -0.002 0.0004 0.0002

∆A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.0000

-0.0002 350

A

50 ns 100 ns 200 ns 500 ns 1 µs 1.5 µs 5 µs

B 400 450 500 550 Wavelength, nm

600

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry

Figure. 3. Nanosecond transient absorption spectra of CuCl42- complexes in acetonitrile (0.1 mMCu(ClO4)2 - 20 mM NEt4Cl, panel A) and dichloromethane (0.1 mM Cu(ClO4)2 - 1 mM NEt4Cl, panel B) after 420-nm excitation. Delay times are shown in the legends.

0.003

∆A

10 mM 20 mM 50 mM 100 mM

465 nm

0.002 0.001 0.000 0.001 0.000

∆A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 46

-0.001 10 mM 20 mM 50 mM 100 mM

-0.002

405 nm

-0.003 0

2000

4000

6000

8000

Time, ns Figure. 4. Nanosecond transient absorption kinetic traces (symbols) of 0.1 mM Cu(ClO4)2 and xmM NEt4Cl (x = 10, 20, 50, and 100 as shown in the legends) acetonitrile solutions measured at probe wavelengths of 465 and 405 nm after 420-nm excitation. Single-exponential fits, ∆A(t) = ∆A0·e-kobst, are shown as lines.

ACS Paragon Plus Environment

12

Page 13 of 46

0.0010

470 nm

1 mM 2 mM 5 mM 20 mM

∆A

0.0005

0.0000

0.0000

∆A

1 mM 2 mM 5 mM

-0.0005

405 nm 0

2000

4000

6000

8000

Time, ns Figure. 5. Nanosecond transient absorption kinetic traces (symbols) of 0.1 mM Cu(ClO4)2 and xmM NEt4Cl (x = 1, 2, 5, and 20) dichloromethane solutions measured at probe wavelengths of 470 and 405 nm after 420-nm excitation. Single-exponential fits, ∆A(t)= ∆A0·e-kobst, are shown as lines. The 405-nm kinetic trace for the 0.1 mM Cu(ClO4)2 - 20 mM NEt4Cl solution is not shown due to significant signals from the scattered excitation light at short times.

7

1.2x10

7

-1

1.0x10

kobs, s

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

8

-1

kq=5.3±0.1*10 s M

-1

6

8.0x10

6

6.0x10

7

-1

kq=9.0±0.4*10 s M

6

-1

4.0x10

6

2.0x10

0.0 0.00

0.02

0.04

0.06

0.08

0.10

[NEt4Cl], M

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 46

Figure. 6. The observed pseudo-first-order rate constants (kobs) plotted as a function of the concentration of NEt4Cl ([NEt4Cl]) in acetonitrile (squares) and dichloromethane (circles). The lines are the linear fits of the kobs values according to the equation: kobs=kq·[NEt4Cl], where the kq are the bimolecular rate constant. The resulting kq values are equal to (9.0 ± 0.1)·107 and (5.3 ± 0.4)·108 M-1 s-1 in acetonitrile and dichloromethane, respectively. Diffusion-controlled rate constants Diffusion-controlled rate constants, kD, for the possible reactions between copper(II) chlorocomplexes were estimated using Smoluchowski relationship for neutral species and Debye theory for ions,27,38 Table 1. The details of these calculations are given in Supporting Information. Table 1. Estimated values of diffusion-controlled rate constants, kD, M-1 s-1 Reaction

Acetonitrile

Dichloromethane

NEt4CuCl3 + NEt4Cl = (NEt4)2CuCl4

N/A

1.6·1010

2NEt4CuCl3 = (NEt4)2Cu2Cl6

N/A

1.6·1010

2CuCl3- = Cu2Cl62-

7.8·109

2.4·109

[CuCl3]- + Cl- = [CuCl4]2-

8.3·109

2.5·109

[CuCl3ACN]- + Cl- = [CuCl4CH3CN]2-

6.5·109

N/A

Calculations The energy profiles for recombination of copper(II) trichlorocomplexes and chloride ions into copper(II) tetrachlorocomplexes are calculated for several reaction mechanisms. The product of ionic dissociation, CuCl3-, exists in acetonitrile as a tetracoorinated solvatocomplex, [CuCl3CH3CN]-, which has D2d local symmetry.20,24 Two reaction mechanisms are proposed for recombination (eq. 1) of a [CuCl3CH3CN]- complex and a Cl- ion in acetonitrile: dissociative (eq.

ACS Paragon Plus Environment

14

Page 15 of 46

2) and associative (eq. 3). Relaxed scans along the Cu−Cl bond of [CuCl4]2- and [CuCl4CH3CN]2- complexes, as well as along the Cu−N bond of [CuCl3CH3CN]- and [CuCl4CH3CN]2- were performed to explore these two mechanisms, Fig. 7 and 8. The resulting values of potential energy barriers, Ea, are shown below. [CuCl3CH3CN]- + Cl- = [CuCl4]2- + CH3CN

(1)

[CuCl3CH3CN]- → [CuCl3]- + CH3CN

Ea = 54.5 kJ mol-1

(2a)

[CuCl3]- + Cl- → [CuCl4]2-

Ea = 3.9 kJ mol-1

(2b)

[CuCl3CH3CN]- + Cl- → [CuCl4CH3CN]2-

Ea = 30.5 kJ mol-1

(3a)

[CuCl4CH3CN]2- → [CuCl4]2- + CH3CN

Ea = 25.3 kJ mol-1

(3b)

M ax

-1

60 50 E, kJ mol

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

-

-

[CuCl3] + Cl + CH3 CN

40 30 20 10 0

2-

-

[CuCl3 CH 3 CN] +Cl

2

3

4

5

[CuCl4] + CH3 CN

-

6

R(Cu-N), Å

7 40 7

6

5

4

3

R(Cu-Cl), Å

Figure. 7. Relaxed redundant coordinate scans along the Cu−N bond of [CuCl3CH3CN]- and along the Cu−Cl bond of [CuCl4]2-in acetonitrile, which explore the dissociative mechanism (eq. 2) of the recombination reaction (eq. 1). The structures corresponding to the [CuCl3CH3CN]-, [CuCl3]-, and [CuCl4]2- energy minima and the barrier top (Max) are shown.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

40 Max2 Max

1

30

-1

20 15 10 5 -

0

25.3 kJ mol

-1

25 30.5 kJ mol

-1

35

E, kJ mol

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 46

2-

[CuCl4CH3CN]

[CuCl3CH3CN] +Cl

40 6

5

4

R(Cu-Cl), Å

2-

[CuCl4] + CH3CN

-

3 2

3

4

5

6 40

R(Cu-N), Å

Figure. 8. Relaxed redundant coordinate scans along the Cu−Cl and Cu−N bonds of the [CuCl4CH3CN]2- complex in acetonitrile exploring the associative mechanism (eq. 3) of the recombination reaction (eq. 1). The structures corresponding to the [CuCl3CH3CN]-, [CuCl4CH3CN]2-, and [CuCl4]2- energy minima and the (Max1 and Max2) energy maxima are shown. We found that in the dissociative mechanism, the rate-liming step is dissociation of [CuCl3CH3CN]- into a [CuCl3]- complex and an acetonitrile molecule (eq. 2a). This step is characterized by the potential energy barrier of 54.5 kJ mol-1. In the associative mechanism, the rate-liming step is the reaction between a [CuCl3CH3CN]- complex and a chloride ion characterized by the potential energy barrier of 30.5 kJ mol-1. This energy barrier is significantly

ACS Paragon Plus Environment

16

Page 17 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

lower than that for the dissociative mechanism. Therefore, according to the DFT calculations, the ligand exchange reaction (eq. 1) proceeds through the associative mechanism (eq. 3). The actual form of a CuCl3- complex is controversial in dichloromethane. We considered three possible forms: [CuCl3]-, [CuCl3CH2Cl2]-, and [Cu2Cl6].19,39 The relaxed scan along the Cu(CuCl3-)−Cl(CH2Cl2) bond (Fig. 9) demonstrates that Cu-CH2Cl2 interaction in the [CuCl3CH2Cl2]- complex is repulsive and, therefore, one can conclude that the [CuCl3CH2Cl2]form of the CuCl3- complex is unfavorable. We did not observe the formation of the CuCl3dimer, [Cu2Cl6]2-, over the about 10-µs time scale investigated. Indeed, for dimerization of free and ion-pair copper(II) trichlorocomplexes, CuCl3- and NEt4CuCl3, diffusion-controlled rate constants, kd are estimated to be 2.4·109 and 1.6 1010 M-1 s-1, respectively, resulting in the fastest reaction time of about 4 and 0.6 µs for the 0.1-mM copper(II) concentration used. However, in the nanosecond transient absorption experiment (Fig. 3), we observed single-exponential decays of the ∆A signals without any spectral shape change up to 8 µs, which is suggestive of relaxation involving a single form of the copper(II) trichlorocomplex. Two other possibilities are that in dichloromethane a copper(II) trichlorocomplex exists as a tricoordinated complex, either CuCl3or NEt4CuCl3. Our calculations demonstrate that the chlorine coordination sphere of CuCl3- is not affected when this complex is approached by NEt4+. Therefore, to simplify our calculations, we performed the relaxed scan along the Cu-Cl bond coordinate of the [CuCl4]2- complex to compute the potential energy profile of the recombination reaction of [CuCl3]- and chloride ions, eq. 4. [CuCl3]- + Cl- → [CuCl4]2-

Ea = 25.0 kJ mol-1

(4)

The potential energy barrier of this recombination reaction is computed to be equal to 25.0 kJ mol-1. Fig. 10 shows the computed energy profile.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry

-

[CuCl3CH2Cl2]

30

E, kJ mol

-1

25 20 15 10 5

-

[CuCl3] + CH2Cl2

0 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5

R(Cu-Cl(CH2Cl2)), Å Figure. 9. The relaxed redundant coordinate scan along the Cu−Cl(CH2Cl2) bond in [CuCl3CH2Cl2]-demonstrates that the Cu-CH2Cl2 interaction is repulsive and, as the result, the [CuCl3CH2Cl2]- form of CuCl3-is unfavorable.

Max

20 10

25.0 kJ mol

-1

E, kJ mol

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 46

0

-

[CuCl3] + Cl

-1

-

-10 -20

2-

[CuCl4] -30

40

35

30

25

20

15

10

5

R(Cu-Cl), Å Figure. 10. The relaxed redundant coordinate scan along the along Cu−Cl bond for [CuCl4]2- in dichloromethane explores the proposed recombination reaction mechanism (eq. 4). The structures corresponding to the energy ([CuCl3]- and [CuCl4]2-) minima and maximum (Max) are shown.

ACS Paragon Plus Environment

18

Page 19 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

IV.

Discussion

Forms of the complexes As shown in the previous studies, a [CuCl4]2- complex has D2d local symmetry in acetonitrile,20,21,36,37 corresponding to the geometry intermediate between square planar and tetrahedral. Because of the significant similarities between absorption spectra of [CuCl4]2- in acetonitrile and dichloromethane (Fig. 1), this complex can be assumed to have D2d local symmetry in dichloromethane as well. Our DFT calculations further support D2d local symmetry of [CuCl4]2- in acetonitrile and dichloromethane (Fig. 8 and 10, Table 3 and 4). Nevertheless, the solvation of [CuCl4]2- is likely to be different in these two solvents. In polar acetonitrile, [CuCl4]2- is solvated by solvent molecules.20,21 In significantly less polar dichloromethane, [CuCl4]2- is surrounded by two tetraethylammonium counter ions and exist in the form of (NEt4+)2(CuCl42-), by analogy with (N(C6H13)4+)2(CuCl42-) in chlorobenzene.40 The ionic dissociation product, CuCl3-, exists in acetonitrile as the tetracoordinated [CuCl3CH3CN]solvatocomplex, which has D2d local symmetry.20,24 A dichloromethane molecule a weak electron donor and, therefore, is unlikely to coordinate to the copper(II) metal center. As a result, the trichlorocopper(II) complex is tricoordinated in dichloromethane. The DFT calculations support this conclusion, Fig. 9. By analogy with [CuCl4]2-, we propose that in the dichloromethane solutions investigated, the CuCl3- complex exists as the (NEt4+)(CuCl3-) ion pair. For these two species, the DFT calculations suggest that the local geometry of the copper(II) ion is trigonal planar. Furthermore, the calculations demonstrate that the geometry of the first coordination sphere is the same for the free CuCl3- and CuCl42- complexes and the (NEt4+)(CuCl3-) and (NEt4+)2(CuCl42-) ion pairs. This is probably why the TDDFT VET energies computed for [CuCl3CH3CN]- and (NEt4+)(CuCl3-) are similar. The level of the TDDFT theory

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 46

used is adequate because the VET energies of [CuCl3CH3CN]- and (NEt4+)(CuCl3-) products agree with the spectral positions of the transient absorption bands, and the VET energies of parent [CuCl4]2- and (NEt4+)2(CuCl42-) tetrachlorocomplexes agree with the steady-state spectra, Fig. 1 and Table 2. Similarly to copper(II) chlorocomplexes, tetraethylammonium chloride species, which are the source of chloride ions, are present as NEt4+Cl- ion pairs in dichloromethane solutions and as fully separated ions in acetonitrile solutions (association constants,Ka = 12 and 92000 M-1in acetonitrile41 and dichloromethane,42 respectively). The complete dissociation of NEt4Cl in acetonitrile is confirmed by the linear dependence of the chloride ion absorbance at 215 nm on the concentration of this salt (Fig. 3S, Supporting information).

Table 2. Energies (in nm) of electronic transitions observed in the steady-state and transient absorption spectra as well as calculated (B3LYP//B3LYP/6-31G(d)/PCM, acetonitrile and dichloromethane parameters) for the copper(II) tri- and tetrachlorocomplexes. The computed oscillator strength values are shown in parentheses. The vertical transitions observed in the experiment are marked in bold. CuCl42-

CuCl42-

(NEt4)2CuCl4

CuCl3CH3CN-

CuCl3-

NEt4CuCl3

in CH3CN

in CH2Cl2

in CH2Cl2

in CH3CN

in CH2Cl2

in CH2Cl2

Calculated

ACS Paragon Plus Environment

20

Page 21 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2428 (0.0003)

2495 (0.0003)

2503 (0.0005)

1572 (0.0006)

2715 (0.0000)

2552 (0.0000)

2428 (0.0003)

2483 (0.0003)

2403 (0.0003)

1483 (0.0004)

1669 (0.0000)

1641 (0.0000)

1977 (0.0003)

2013 (0.0003)

2002(0.0003)

1173 (0.0009)

1573 (0.0000)

1497 (0.0000)

1391 (0.0000)

1405 (0.0000)

1386 (0.0000)

1050 (0.0000)

1421 (0.0003)

1371 (0.0003)

442 (0.0000)

443 (0.0000)

449 (0.0001)

487 (0.0064)

463 (0.0222)

465 (0.0249)

419 (0.0141)

420 (0.0143)

433 (0.0093)

435 (0.0089)

401 (0.0000)

399 (0.0001)

419 (0.0142)

420 (0.0144)

426 (0.0207)

396 (0.0178)

380 (0.0002)

385 (0.0002)

360 (0.0039)

360 (0.0042)

371 (0.0028)

381 (0.0016)

374 (0.0000)

380 (0.0000)

341 (0.0001)

341 (0.0003)

357 (0.0013)

365 (0.0012)

357 (0.0119)

366 (0.0096)

340 (0.0002)

341 (0.0002)

350 (0.0008)

344 (0.0005)

343 (0.0384)

343 (0.0360)

338 (0.0000)

338 (0.0000)

344 (0.0005)

314 (0.0755)

306 (0.0172)

311 (0.0166)

318 (0.0000)

319 (0.0000)

331 (0.0002)

292 (0.0560)

277 (0.0327)

276 (0.0266)

298 (0.0544)

298 (0.0554)

307 (0.0590)

260 (0.0325)

248 (0.0581)

250 (0.0601)

297 (0.0544)

297 (0.0551)

298 (0.0684)

273 (0.0000)

273 (0.0000)

277 (0.0004)

241 (0.0414)

241 (0.0433)

244 (0.0326)

Experimental 460 (shoulder)

460 (shoulder)

350

350

407

411

465

470

345 (shoulder)

345 (shoulder)

294

295

241

241

Reaction mechanism In the numbers of studies,32,33,40,43,44 it was shown by means of steady-state photolysis, actinometry, and transient absorption techniques that important relaxation channels of the LMCT excited states of [CuCl4]2- complexes are internal conversion to the ground state through the

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 46

manifold of lower-lying ligand-field excited states, ionic dissociation into a [CuCl3]- complex and a Cl- ion, and radical dissociation into a [CuCl3]2- complex and a chlorine atom. The ionic dissociation is the major reaction channel for [CuCl4]2- complexes in the LMCT excited states. Indeed, quantum yields of the formation of copper(I) complexes, products of the radical dissociation do not exceed 0.5%32,44 and are significantly smaller than quantum yields of the formation of the ionic dissociation products. However, due to high lability intrinsic to copper(II) complexes the ionic dissociation products quickly recombine to reform the parent complexes, whereas the radical dissociation products undergo further reactions forming long-lived photoreduction products, which include stable copper(I) complexes. Excitation at 420 nm predominantly causes the 2ECl(n) → Cu(dxy) LMCT transition in the [CuCl4]2complex.33 The populated Franck-Condon LMCT state is repulsive, leading to the Cu-Cl bond breaking and the formation of copper(II) trichlorocomplex and chloride ion products.32,33 In the previous studies, the 20 ps - 1 ns transient absorption spectra of [CuCl4]2- in acetonitrile were assigned to the formation of the [CuCl3CH3CN]- product of ionic dissociation (eq. 5, 6):32,33 420nm

[Cu IICl4 ]2 −  → [Cu IICl3 ]− + Cl−

(5)

< 20ps

[Cu IICl3 ]− + CH3CN → [Cu IICl3CH3CN]−

(6)

The lack of change of the ∆A spectra after 20 ps demonstrates that the initially formed [CuCl3]complexes react with the acetonitrile solvent molecules well within the first 20 ps after excitation, forming the [CuCl3CH3CN]- complexes. This conclusion agrees with the DFT calculations, which demonstrate that the reaction between [CuCl3]- and acetonitrile has no potential energy barrier, Fig. 7. The exact built-up time constant of [CuCl3CH3CN]- cannot be determined because the short-time ∆A spectra are dominated by intense ligand-field excited-state absorption.32,33 The 20-ps - 1-ns ∆A spectra of dichloromethane solutions are similar to those in

ACS Paragon Plus Environment

22

Page 23 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

acetonitrile solutions, albeit with a small solvatochromic shift of the 465-nm band to 470 nm, and can analogously be assigned to ligand loss by (NEt4+)2(CuCl42-) with the formation of (NEt4+)(CuCl3-). Based on the DFT calculations, the formation of (NEt4+)(CuCl3CH2Cl2-) complexes is unlikely. The spectral shape of the nanosecond ∆A spectra is similar to the picosecond ∆A spectra in both acetonitrile and dichloromethane, Fig. 3. This observation is evidence that starting from 20 ps up to several µs, the product copper(II) species exist in the single form in these two solvents, [CuCl3CH3CN]- and (NEt4+)(CuCl3-). The positive ∆A bands decays single-exponentially to the zero signal level concurrently with the ground-state bleach recovery and without spectral shape change, revealing a single process, namely, the recombination reaction of copper(II) trichlorocomplexes and chloride ions into parent copper(II) tetrachlorocomplexes (eq. 1, 7): [NEt4][CuCl3] + NEt4Cl= [NEt4]2 [CuCl4]

(7)

The recombination pseudo-first-order rate constants linearly depend on the concentration of NEt4Cl, Fig. 6: kobs=kq·[NEt4Cl]

(8)

The bimolecular rate constants, kq, are equal to (9.0 ± 0.1)·107 and (5.3 ± 0.4)·108 M-1 s-1 in acetonitrile and dichloromethane, respectively. These values are about two order smaller than the diffusion-controlled rate constants, 8.3·109 and 1.6·1010M-1 s-1 estimated in acetonitrile and dichloromethane, respectively, Table 1, implying that the corresponding recombination reactions in the both solvents are controlled by potential energy barriers. In acetonitrile, the recombination reaction of copper(II) trichlorocomplexes and chloride ions to parent[CuCl4]2- complexes, eq. (1), occurs via ligand exchange. The large bimolecular rate constant of this ligand exchange, 9.0·107 M-1 s-1, demonstrates that the copper(II)

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 46

chlorocomplexes are highly labile in acetonitrile.7 The observation that the ligand recombination rate constant linearly depends on the concentration of the incoming ligand, the chloride ions, implies that the ligand exchange reaction (eq. 1) proceeds either through the associative or associative interchange mechanisms (eq. 3). In our experiments, only the single form of copper(II) trichlorocomplex is observed whereas intermediate species are not kinetically detected, which leads to the conclusion that ligand exchange between a [CuCl3CH3CN]- complex and a Cl- ion is described by the associative interchange mechanism. The proposed mechanism is supported by the DFT calculations. In the both associative and dissociative mechanisms (eq. 2, 3), the first steps, which, correspondingly, are addition of the chloride ion and elimination of the acetonitrile molecule, are rate-limiting as they are controlled by large potential energy barriers (30.5 and 54.5 kJ mol-1, respectively, Fig. 7 and 8). As a result, for the ligand exchange reaction between a [CuCl3CH3CN]- complex and a Cl- ion the associative mechanism is favorable. In dichloromethane, recombination of copper(II) trichlorocomplexes and chloride ions to parent copper(II) tetrachlorocomplexes, eq. (4) is a simple one-step addition reaction. According to the DFT calculations, the potential energy barrier for this reaction is 25.0 kJ mol-1, Fig. 10. Thus, the potential energy barriers for recombination of copper(II) trichlorocomplexes and chloride ions into copper(II) tetrachlorocomplexes are 30.5 and 25.0 kJ mol-1 in acetonitrile and dichloromethane, respectively. This trend is in agreement with the experimental recombination rate constants being larger in dichloromethane (5.3·108 M-1 s-1) than in acetonitrile (9.0·107 M-1 s-1) because lowering the energy barrier typically accelerates the reaction. Faster recombination in acetonitrile than in dichloromethane can be explained by steric/geometrical factors. The addition of a chloride ion to a tricoordinated[CuCl3]- in dichloromethane (eq. 4) does not seem to involve significant geometry rearrangement of the reactant species because at the energy

ACS Paragon Plus Environment

24

Page 25 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

maximum, the CuCl3 moiety has the same geometry as in the initial [CuCl3]- complex. Therefore, the potential energy barrier of the addition arises as the result of columbic repulsion between two negatively charged ions, Fig 10 and Table 3. At the same time, the potential energy barrier for the addition of a chloride ion to a tetracoordinated[CuCl3CH3CN]- in acetonitrile (eq. 3a) is due to both columbic repulsion between two negatively charged ions and steric factors because in the maximum energy geometry, the CuCl3CH3CN fragment has the geometry significantly different from that of the initial [CuCl3CH3CN]- complex, Table 4. This implies that additional energy is required for the addition in acetonitrile.

Table 3. Structural parameters of the optimized geometries of the reagents, products, and potential energy maximum of the recombination reaction between a copper(II) trichlorocomplex and a chloride ion to form a copper(II) tetrachlorocomplex in dichloromethane. All calculations were performed at the B3LYP/6-31G(d)//PCM(dichloromethane) level of theory. [CuCl3]-

Max1

[CuCl4]2-

Cu−Cl1

2.19

2.19

2.30

Cu−Cl2

2.20

2.21

2.31

Cu−Cl3

2.19

2.19

2.30

4.00

2.31

Bond length, Å Cu−Cl4

Angle, degrees

Cl1−Cu−Cl2

112.5

110.7

101.0

Cl2−Cu−Cl3

111.8

110.8

101.0

Cl1−Cu−Cl3

135.7

136.4

128.0

Cl1−Cu−Cl4

89.5

101.1

Cl2−Cu−Cl4

108.5

128.2

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 46

Table 4. Structural parameters of the optimized geometries of the reagents, products, and potential energy maxima of the recombination reaction between a copper(II) trichlorocomplex and a chloride ion to form a copper(II) tetrachlorocomplex in acetonitrile. All calculations were performed at the B3LYP/6-31G(d)//PCM(dichloromethane) level of theory.

Bond length, Å

[CuCl3CH3CN]-

Max1

[CuCl4CH3CN]2- Max2

[CuCl4]2-

Cu−Cl1

2.28

2.36

2.34

2.33

2.31

Cu−Cl2

2.28

2.33

2.33

2.32

2.31

Cu−Cl3

2.28

2.43

2.50

2.32

2.31

3.20

2.50

2.48

2.31

Cu−Cl4

Angle, degrees

Cu−N

1.96

1.97

1.96

2.66

Cl1−Cu−Cl2

100.4

92.2

96.7

95.7

100.8

Cl2−Cu−Cl3

100.4

92.6

116.6

135.5

100.9

Cl1−Cu−Cl3

137.2

171.7

91.4

95.6

128.7

Cl1−Cu−Cl4

89.6

91.1

103.3

100.8

Cl2−Cu−Cl4

107.0

119

108.8

128.5

Cl3−Cu−Cl4

95.6

123.5

110

100.9

Cl1−Cu−N

95.3

88.3

177.4

172.1

Cl2−Cu−N

135.6

158.1

86.0

83.1

Cl3−Cu−N

46.3

84.8

87.5

80.0

94.9

87.5

84.5

Cl4−Cu−N

V.

Conclusions

The dynamics and underlying reaction mechanisms of the formation of CuCl42- complexes from copper(II) CuCl3- trichlorocomplexes and chloride ions, the recombining products of ultrafast dissociation of CuCl42- promoted into the LMCT excited state upon 420-nm excitation,

ACS Paragon Plus Environment

26

Page 27 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

are studied in two solvents with different donor properties (acetonitrile and dichloromethane) by means of UV-vis transient absorption when probing from 20 ps to 8 µs, in conjunction with DFT/TDDFT calculations. In acetonitrile solutions, the CuCl3- products rapidly reacts with the solvent molecule in less than 20 ps to form the [CuCl3CH3CN]-solvatocomplexes, which then recombine with chloride ions to reform parent [CuCl4]2- complexes. In dichloromethane solutions, CuCl3- products do not form solvatocomplexes due to weak donor properties of CH2Cl2 molecules. Instead, they immediately recombine with chloride ions to reform parent [CuCl4]2- complexes. It is important to note, that all negatively charged species exist as solvated ions in acetonitrile but they form ion pairs with NEt4+ counter ions in dichloromethane because of low polarity of this solvent. In the both solvents investigated, the recombination rate constants linearly depend on the concentration of NEt4Cl, the source of chloride-ions. The bimolecular rate constants of the recombination, kq, are equal to (9.0 ± 0.1)·107 and (5.3 ± 0.4)·108 M-1 s-1 in acetonitrile and dichloromethane, respectively, which are about two orders of magnitude smaller than diffusion-controlled rate constants in these solvents, implying that in the both cases the recombination reactions are controlled by energy barriers. In

acetonitrile,

the recombination

reaction

between

copper(II)

trichlorocomplexes,

[CuCl3CH3CN]-, and chloride ions to reform [CuCl4]2- complexes occurs via ligand exchange, which can be interpreted as the associative interchange mechanism with the intermediacy of pentacoordinated complexes, [CuCl4CH3CN]2-: −

Cl [CuCl3CH3CN]- + → [CuCl4 CH3CN]2- −  → [CuCl4 ]2CH3CN

(9)

The proposed mechanism is supported by the DFT calculations. The calculated value of the activation energy for the rate-limiting step, which is the addition of Cl-, is about 30.5 kJ mol-1. In

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 46

dichloromethane, the recombination of copper(II) trichlorocomplexes and chloride ions into copper(II) tetrachlorocomplexes, takes place via a simple one-step addition: [CuCl3]- + Cl- → [CuCl4]2-

(10)

According to our DFT calculations, the potential energy barrier for this reaction is 25.0 kJ mol1

. The recombination in acetonitrile requires a larger energy barrier compared with that in

dichloromethane due to steric/geometrical factors. This trend correlates to that for the recombination bimolecular rate constants, which are observed to be larger in dichloromethane than in acetonitrile.

Supporting Information Supporting Information section contains 1) detailed description of the calculation of diffusioncontrolled rate constants, 2) steady-state spectra of copper(II) solutions in acetonitrile and dichloromethane used in transient absorption measurements and tetraethylammonium chloride solutions in acetonitrile, and 3) cartesian coordinates of the optimized geometries of copper(II) chlorocomplexes and corresponding energies of potential-energy minima and maxima. This information is available free of charge via the Internet at http://pubs.acs.org Corresponding Author *To whom correspondence should be addressed. Address: Institute of Chemistry, SaintPetersburg State University, 26 Universitetskiy pr., Petrodvorets, 198504 Saint-Petersburg, Russian Federation; E-mail: [email protected]; Phone: +7-951-677-5465 Author Contributions

ACS Paragon Plus Environment

28

Page 29 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding Sources Saint-Petersburg State University, National Science Foundation, Russian Foundation for Basic Research, Ohio Supercomputer Center.

Acknowledgments A.S.M. and M.S.P. acknowledge the Saint-Petersburg State University for the financial support (postdoctoral fellowships No. 12.50.1562.2013 and 12.50.1189.2014). M.Yu.S. acknowledges Saint-Petersburg State University for the financial support in framework of Event 6, grant 12.42.1270.2014. Optical measurements were performed at the Center for Optical and Laser Materials Research of the Saint-Petersburg State University and the Center for Photochemical Sciences, Bowling Green State University. Calculations were carried out at the Saint-Petersburg State University Computer Center and the Ohio Supercomputer Center (PCS0204-7). This work was supported by the NSF CAREER (CHE-0847707, A.N.T.), NSF MRI (CHE-0923360, A.N.T.), Saint Petersburg State University research (2015–2017, 12.38.219.2015), and RFBR (14-03-01003 and 15-03-05139) awards.

References (1)

Hervas, M.; Navarro J. A.; Diaz, A.; Bottin, H.; De la Rosa, M.A. Laser-Flash Kinetic

Analysis of the Fast Electron Transfer from Plastocyanin and Cytochrome C6 to Photosystem I. Experimental Evidence on the Evolution of the Reaction Mechanism. Biochem. 1995, 34, 1132111326.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(2)

Page 30 of 46

Rasmussen, M.; Minteer, S. D.; Investigating the Mechanism of Thylakoid Direct

Electron Transfer Forphotocurrent Generation. Electrochim.Acta 2014, 126, 68–73. (3)

Ranieri, A.; Battistuzzi, G.; Borsari, M.; Casalini, S.; Fontanesi, C.; Monari, S.; Siwek,

M. J.; Sola, M. Thermodynamics and Kinetics of the Electron Transfer Process of Spinach Plastocyanin Adsorbed on a Modified Gold Electrode. J. of Electroanal. Chem. 2009, 626, 123– 129. (4)

Olloqui-Sariego, J. L.; Moreno-Beltran, B.; Diaz-Quintana, A.; De la Rosa, M.; Calvente,

J. J.; Andreu, R. Temperature-Driven Changeover in the Electron-Transfer Mechanism of a Thermophilic Plastocyanin. J. Phys. Chem. Lett. 2014, 5, 910−914. (5)

Solomon, E. I.; Szilagyi, R.K.; George, S. D.; Basumallick, L. Electronic Structures of

Metal Sites in Proteins and Models: Contributions to Function in Blue Copper Proteins. Chem. Rev. 2004, 104, 419-458. (6)

Nagasawa, Y.; Fujita, K.; Katayama, T.; Ishibashi, Y.; Miyasaka, H.; Takabe, T.; Nagao,

S.; Hirota, S. Coherent Dynamics and Ultrafast Excited State Relaxation of Blue Copper Protein; Plastocyanin. Phys. Chem. Chem. Phys. 2010, 12, 6067–6075. (7)

Richens, D. T. Ligand Substitution Reactions at Inorganic Centers. Chem. Rev. 2005,

105, 1961–2002. (8)

Vafazadeh, R.; Bidaki, S.Kinetics and Mechanism of Ligand Exchange Reaction of

Copper(II) Complexes with Tetradentate Schiff Base Ligands. Acta Chim. Slov. 2014, 61, 153– 160. (9)

Helm, L.; Merbach, A. E. Inorganic and Bioinorganic Solvent Exchange Mechanisms.

Chem. Rev. 2005, 105, 1923–1959.

ACS Paragon Plus Environment

30

Page 31 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(10) Cui, X.; Khlobystov, A. N.; Chen, X.; Marsh, D. H.; Blake, A. J.; Lewis, W.; Champness, N. R.; Roberts, C. J.; Schröder, M. Dynamic Equilibria in Solvent-Mediated Anion, Cation and Ligand Exchange in Transition-Metal Coordination Polymers: Solid-State Transfer or Recrystallisation. Chem. Eur. J. 2009, 15, 8861 – 8873. (11) Rotzinger, F. P. Treatment of Substitution and Rearrangement Mechanisms of Transition Metal Complexes with Quantum Chemical Methods. Chem. Rev. 2005, 105, 2003−2037. (12) Maigut, J.; Meier, R.; Zahi, A.; van Eldik, R. Effect of Chelate Dynamics on Water Exchange Reactions of Paramagnetic Aminopolycarboxylate Complexes. Inorg. Chem. 2008, 47, 5702-5719. (13) Ma, H.; Wan, C.; Zewail, A. H. Dynamics of Ligand Substitution in Labile Cobalt Complexes Resolved by Ultrafast T-Jump. Proc. Natl. Acad. Sci. USA, 2008, 105, 12754-12757. (14) Schwenk, C. F.; Loeffler, H. H.; Rode B. M. Structure and Dynamics of Metal Ions in Solution:  QM/MM Molecular Dynamics Simulations of Mn2+ and V2+. J. Am. Chem. Soc., 2003, 125, 1618-1624. (15) Kawamura, K.; Igarashi, S.; Yotsuyanagi, T. Kinetics of the Accelerator Ligand Effect for the Complex Formation of Cd(II) Ion with a Cationic Water-Soluble Porphyrin. Inorg. Chim. Acta 2007, 360, 3287-3295. (16) Rode, B. M.; Shwenk, C. F.; Hofer, T. S.; Randolf, B. R. Coordination and Ligand Exchange Dynamics of Solvated Metal Ions. Coord. Chem. Rev. 2005, 349, 2993-3006. (17) Vos, J. G.; Pryce, M. T. Photoinduced Rearrangements in Transition Metal Compounds. Coord. Chem. Rev. 2010, 254, 2519-2531. (18) Taube, H. Rates and Mechanisms of Substitution in Inorganic Complexes in Solution. Chem. Rev. 1952, 50, 69–126.

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 46

(19) Langford, C. H.; Gray, H. B. Ligand Substitution Processes; W. A. Benjamin, Inc.: New York, 1965. (20) Ohtaki, H. Structural Studies on Solvation and Complexation of Metal Ions in Nonaqueous Solutions. Pure Appl. Chem. 1987, 59, 1143–1150. (21) Ishiguro, S.; Jeliazkova, B. G.; Ohtaki, H. Complex Formation and Solvation of (CuCln)(2-n)+ in Acetonitrile and in N,N-dimethylformamide. Bull. Chem. Soc. Jpn. 1985, 58, 1749–1754. (22) Khan, M.; Bouet, G.; Vierling, F.; Meullemeestre, J.; Schwing, M.-J. Formation of Cobalt(II), Nickel(II) and Copper(II) Chloro Complexes in Alcohols and the Irving-Williams Order of Stabilities. Transit. Met. Chem. 1996, 21, 231–234. (23) Khan,

M.

A.;

Meullemeestre,

J.;

Schwing,

M.

J.;

Vierling,

F.

Detailed

Spectrophotometric Study of the Copper(II) Halides in Anhydrous Methanol. Inorg. Chem. 1989, 7, 3306–3309. (24) Kudrev, A. G. Calculation of the Equilibrium Constants and Cooperative Parameters of Formation of Cu(II) Chloride Complexes in Nonaqueous Solvents. Russ. J. Coord. Chem. 2010, 36, 704–710. (25) Helm,

L.;

Lincoln,

I.

S.

F.;

Zbindenla,

D.

Solvent

Exchange

in

Hexakis(methanol)copper(II) Ion. Oxygen-17 NMR Variable-Temperature, Pressure, and Frequency Study. Inorg. Chem. 1986, 25, 2550–2552. (26) Eigen, M.; Wikinns, R. G. The Kinetics and Mechanism of Formation of Metal Complexes. Adv. Chem. Ser. 1965, 49, 55–80.

ACS Paragon Plus Environment

32

Page 33 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(27) Mereshchenko, A. S.; Pal, S. K.; Karabaeva, K. E.; El-Khoury, P. Z.; Tarnovsky, A. N. Photochemistry of Monochloro Complexes of Copper(II) in Methanol Probed by Ultrafast Transient Absorption Spectroscopy. J. Phys. Chem. A 2012, 116, 2791–2799. (28) Maigut, J.; Meier, R.; Zahi, A.; van Eldik, R. Triggering Water Exchange Mechanisms via Chelate Architecture. Shielding of Transition Metal Centers by Aminopolycarboxylate Spectator Ligands. J. Am. Chem. Soc. 2008, 130, 14556–14569. (29) Houston, J. R.; Richens, D. T.; Casey, W. H. Distinct Water-Exchange Mechanisms for Trinuclear Transition-Metal Clusters. Inorg. Chem. 2006, 45, 7962-7967. (30) Caër, S. L.; Mestdagh, M.; Maître, P. Condensation Reaction vs. Ligand Exchange with First-Row Transition Metal Cations: a Theoretical Study of Cu+ Heteroleptic Model Complexes. Int. J. Mass. Spectrom. 2003, 227, 587–600. (31) Ma, H. R.; Wan, C. Z.; Zewail, A. H. Ultrafast T-Jump in Water: Studies of Conformation and Reaction Dynamics at the Thermal Limit. J. Am. Chem. Soc. 2006, 128, 6338–6340. (32) Mereshchenko, A. S.; Olshin, P. K.; Karimov, A. M.; Yu, M.; Burkov, K. A.; Tveryanovich, Y. S.; Tarnovsky, A. N. Photochemistry of Copper(II) Chlorocomplexes in Acetonitrile: Trapping the Ligand-to-Metal Charge Transfer Excited State Relaxations Pathways. Chem. Phys. Lett. 2014, 615, 105–110. (33) Mereshchenko, A. S. Ultrafast Photochemistry of Polyatomic Molecules Containing Lbile Halogen Atoms in Solution. Ph.D. Thesis, Bowling Green State University, Bowling Green, OH, 2013. (34) CRC Handbook of Chemistry and Physics; Lide, D. R., Ed.; 90th ed.; CRC Press (Taylor and Francis Group): Boca Raton, FL, 2009. (35) Frisch, M. J. et al. Gaussian 09, Revision A.02. Gaussian 09, Revision A.02, 2009.

ACS Paragon Plus Environment

33

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 46

(36) Ferguson, J. Electronic Absorption Spectrum and Structure of CuCl4=. J. Chem. Phys. 1964, 40, 3406–3410. (37) Furlani, C.; Morpurgo, G. Properties and Electronic Structure of Tetrahalogenocuprate (II)-Complexes. Theor. Chim. Acta 1963, 1, 102–115. (38) Rice, S. A. Diffusion-Limited Reactions; Bamford, C. H.; Tipper, C. F. H.; Compton, R. G., Eds.; Comprehensive Chemical Kinetics; Elsevier, 1985; Vol. 25, pp. 3–46. (39) Doyle, K. J.; Tran, H.; Baldoni-Olivencia, M.; Karabulut, M.; Hoggard, P. E. Photocatalytic Degradation of Dichloromethane by Chlorocuprate(II) Ions. Inorg. Chem. 2008, 47, 7029–7034. (40) Golubeva, E. N.; Zubanova, E. M.; Melnikov, M. Y.; Gostev, F. E.; Shelaev, V.; Nadtochenko, V. A. Femtosecond Spectroscopy and TD-DFT Calculations of CuCl4 2− Excited States. Dalt. Trans. 2014, 43, 17820–17827. (41) Grimsrud, E. P.; Kratochvil, B. Second-Order Rate Constants and Ion Association in Aprotic Solvents by High-Precision Conductance. J. Am. Chem. Soc. 1973, 95, 4477–4483. (42) Bait, S.; du Chattel, G.; de Kieviet, W.; Tieleman, A. Ion Association and Solvation in Dichloromethane of Tetrachloro- and Tetrabromoferrates(III) Compared with Simple Halides. Naturforsch B, 1978, 33, 745–749. (43) Gromov, O. I.; Zubanova, E. M.; Golubeva, E. N.; Plyusnin, V. F.; Zhidomirov, G. M.; Melnikov, M. Y. UV–Vis Identification and DFT-Assisted Prediction of Structures of Cu(II)– Alkyl Chlorocomplexes. J. Phys. Chem. A 2012, 116, 11581–11585. (44) Cervone, E.; Diomedi Camassei, F.; Giannini, I.; Sykora, J. Photoredox Behaviour of Chlorocopper(II) Complexes in Acetonitrile: Mechanism and Quantum Yields. J. Photochem. 1979, 11, 321–332.

ACS Paragon Plus Environment

34

Page 35 of 46

-

6

lorom

10

etha ne k =5 q .3*10 8 -1 -1 s M

12

8

-

[CuCl3] + Cl

Dich

6 -1

Table of Contents (TOC) Image

Recombination rate constant, 10 s

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

4

[CuCl4]

2-

-

krecomb=kq*[Cl ]

Ac

rile nit o t e

9 k q=

7

1 .0*

0

-1 -1

s

M

2 0 0.00

0.02

0.04 0.06 [Cl ], M

0.08

0.10

ACS Paragon Plus Environment

35

The Journal of Physical Chemistry

-1

6000

4000

-1

, M cm

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 46

2000

0

250 300 350 400 450 500 550 600

, nm  

Figure. 1. Mereshchenko A.S. et al. ACS Paragon Plus Environment

Page 37 of 46

0.0004

A

0.0000 20 ps 50 ps 100 ps 200 ps 500 ps 1 ns

A

-0.0004 -0.0008 0.0004

A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.0000

20 ps 50 ps 100 ps 200 ps 500 ps 1 ns

B

-0.0004 350

400

450

500

550

Wavelength, nm

Figure. 2. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

600

The Journal of Physical Chemistry

0.002

A

0.001 0.000

50 ns 100 ns 200 ns 500 ns 1 s 2 s 5 s

-0.001 -0.002 0.0004

0.0000

B

-0.0002 350

A

50 ns 100 ns 200 ns 500 ns 1 s 1.5 s 5 s

0.0002

A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 46

400 450 500 550 Wavelength, nm

Figure. 3. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

600

Page 39 of 46

0.003

A

10 mM 20 mM 50 mM 100 mM

465 nm

0.002 0.001 0.000 0.001 0.000

A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

-0.001 10 mM 20 mM 50 mM 100 mM

-0.002

405 nm

-0.003 0

2000

4000

6000

Time, ns

Figure. 4. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

8000

The Journal of Physical Chemistry

0.0010

470 nm

1 mM 2 mM 5 mM 20 mM

A

0.0005

0.0000

0.0000 1 mM 2 mM 5 mM

A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 46

-0.0005

405 nm 0

2000

4000

6000

Time, ns

Figure. 5. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

8000

Page 41 of 46

7

1.2x10

8

7

-1

-1

-1

kq=5.3±0.1*10 s M

1.0x10

kobs, s

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

6

8.0x10

6

6.0x10

7

6

4.0x10

6

2.0x10

0.0 0.00

-1

-1

kq=9.0±0.4*10 s M

0.02

0.04

0.06

0.08

[NEt4Cl], M

Figure. 6. Mereshchenko A.S. et al. ACS Paragon Plus Environment

0.10

The Journal of Physical Chemistry

M ax

-1

60 50 E, kJ mol

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 42 of 46

-

-

[CuCl3] + Cl + CH3 CN

40 30 20 10 0

2-

-

[CuCl3 CH 3 CN] +Cl

2

3

4

5

[CuCl4] + CH3 CN

-

6

R(Cu-N), Å

7 40 7

6

5

4

R(Cu-Cl), Å

Figure. 7. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

3

Page 43 of 46

40 Max2 Max

1

30

-1

20 15 10 5 -

0

25.3 kJ mol

-1

25 30.5 kJ mol

-1

35

E, kJ mol

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2-

[CuCl4CH3CN]

[CuCl3CH3CN] +Cl

40 6

5

4

R(Cu-Cl), Å

2-

[CuCl4] + CH3CN

-

3 2

3

4

5

R(Cu-N), Å

Figure. 8. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

6 40

The Journal of Physical Chemistry

30

-

[CuCl3CH2Cl2]

-1

25

E, kJ mol

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 46

20 15 10 5

-

[CuCl3] + CH2Cl2

0 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5

R(Cu-Cl(CH2Cl2)), Å

Figure. 9. Mereshchenko A.S. et al. ACS Paragon Plus Environment

Page 45 of 46

Max

20 10

25.0 kJ mol

E, kJ mol

-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0

-

[CuCl3] + Cl

-1

-

-10 -20 [CuCl4] -30

40

35

30

25

20

15

R(Cu-Cl), Å

Figure. 10. Mereshchenko A.S. et al.

ACS Paragon Plus Environment

10

5

2-

The Journal of Physical Chemistry

-

2-

4

[CuCl4] -

krecomb=kq*[Cl ]

etha lorom

10

6

ne k =5 q .3*10 8 -1 s M -1

12

8

-

[CuCl3] + Cl

D ic h

6 -1

Table of Contents (TOC) Image

Recombination rate constant, 10 s

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 46

A

to ce

9 k q=

.

r ni t

ile

-1 -1

7

0 0 *1

s

M

2 0 0.00

0.02

0.04 0.06 [Cl ], M

0.08

ACS Paragon Plus Environment

0.10